You are on page 1of 15

Journal of Hydrology 235 (2000) 12–26

www.elsevier.com/locate/jhydrol

Estimation of hydraulic parameters of shaly sandstone aquifers


from geoelectrical measurements
O.A.L. de Lima a,*, Sri Niwas b
a
Center for Research in Geophysics and Geology (CPGG)-Federal University of Bahia, Campus Universitário da Federação,
Salvador, Bahia 40170-290, Brazil
b
Department of Earth Sciences, University of Roorkee, Roorkee 247 667, India
Received 5 October 1999; revised 7 March 2000; accepted 6 April 2000

Abstract
Electrical and hydraulic conductivities of shaly sandstones are described using a capillary approach for a granular, clay
bearing material. The clays are assumed to occur as shale shells uniformly coating the insulated sand grains. The real and
imaginary components of the complex electrical conductivity for this material model are written in terms of a unique solid
matrix conductivity, treated as a volumetric property. Above a critical salt concentration the conductive contribution of the
shells is independent of the electrolyte salinity. Under this condition the equation for the bulk conductivity of the sandstone can
be expressed in a simplified form. However, below the critical concentration the matrix conductivity is dependent on the
conductivity of the water saturating the shaly component and can be expressed only by the complete equation. The hydraulic
conductivity for this model is expressed by a modified Kozeny–Carman equation. A new semi-empirical equation relates the
hydraulic conductivity for such rocks, to their porosity, formation resistivity factor and the electrical conductivity of its solid
matrix. These combined properties are described as the lithoporosity factor. In this new formulation the petrophysical para-
meters involved are easily determined from the electrical geophysical measurements. The performance of this equation is firmly
tested with experimental laboratory data available in the literature. Its application is then extended to estimate the hydraulic
parameters of a shaly sandstone aquifer in Bahia-Brazil, using either the borehole or the surface geoelectrical data. Examples
are given to emphasize the combined use of electrical resistivity and induced polarization measurements in computing
hydraulic properties. 䉷 2000 Elsevier Science B.V. All rights reserved.
Keywords: Permeability; Electrical conductivity; Induced polarization; Shaly sand-stones; Geophysical logging

1. Introduction aquifer behavior one needs to solve a set of differen-


tial equations under appropriate boundary conditions.
The knowledge of the spatial distribution of In the absence of exact solution of these equations,
hydraulic parameters of a groundwater flow domain numerical solutions are sought by subdividing the
is essential both for optimizing aquifer exploration aquifer in many blocks of elementary dimensions.
programs as well as to evaluate the efficiency of For heterogeneous aquifers the procedure of assigning
contaminant remediation processes. To model the the hydraulic parameters of each element is not
obvious.
* Corresponding author.
Analysis of pump test data as a function of time are
E-mail addresses: olivar@cpgg.ufba.br (O.A.L. de Lima), classical procedures for estimating aquifer para-
srsnpfes@rurkiu.ernet.in (Sri Niwas). meters. These results suffer from non-uniqueness
0022-1694/00/$ - see front matter 䉷 2000 Elsevier Science B.V. All rights reserved.
PII: S0022-169 4(00)00256-0
O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26 13

and because of physical and conceptual constraints where a and q are empirical constants and F, the
underlying such tests, the interpreted results represent formation resistivity factor, is given as F ˆ s w =s 0 ;
macroscopic averages taken over large volumes of the with s 0 and s w being the conductivities of the fully
porous medium. Small-scale textural and structural saturated whole rock and of its pore electrolyte,
variabilities are lost in this process and these details respectively. The constant a in Eq. (1) is inversely
may be necessary to solve the hydrological problem proportional to the square of the specific surface for
efficiently. the sandstone, and so has a dimension of L 2.
Geophysical measurements made within wells or at It is worthwhile to recall here that Heigold et al.
Earth’s surface sample small formation volumes, typi- (1980) have combined Darcy’s equation and Archie’s
cally within a decimetric to a decametric scale. The relation (Archie (1950)) between permeability and
physical conditions (tortuosity and porosity) control- porosity of rocks to obtain the following equation
ling the electric current flow (electrical resistivity) similar to Eq. (1) as
also likewise control the lateral flow of the water
(hydraulic conductivity) in porous media. Exploiting k ˆ a 1 fb1 ; …2†
this similarity, a large number of empirical equations where k is the intrinsic permeability (m 2)—the pore
has been developed to convert the geophysical area of the rock governing the flow, f is porosity and
measurements into hydraulic parameters (Kelly, a1 and b1 are constants. However, the hydrophysical
1977; Kosinski and Kelly, 1981; Scott, 1988; Sen et property is the hydraulic conductivity, K (m/s)—fluid
al., 1990; Börner et al., 1996; Sighal et al., 1998; velocity, which depends on both the formation and the
Yadav and Abolfazli, 1998). Sri Niwas and Singhal fluid contained in it. Notwithstanding, our effort
(1981, 1985) have developed an analytical relation- should be to obtain a more physically supported
ship between the hydraulic conductivity and electrical quantitative relation instead of an empirical one.
resistivity through the well-established laws of Nuttings’ equation (Hubbert, 1940) relates k and K
physics, i.e. Darcy’s law of lateral flow of ground as
water and Ohm’s law of current flow in clean porous rg
media. These results provide a physical and mathema- Kˆk ; …3†
m
tical basis to the statistically established relations.
For clean sandstones, a porous model of free rugose where m is the dynamic viscosity of the fluid (kg/ms),
channels leads to a fractal representation for its poros- r the fluid density (kg/m 3) and g the acceleration due
ity, specific surface, hydraulic and electrical tortuos- to gravity (m/s 2). For hydrological problems the fluid
ities as well as to its formation resistivity factor (Katz is water and ideally m and r can be taken as unit. Thus
and Thompson, 1985; Korvin, 1992; Turcotte, 1992). the permeability value can be converted approxi-
In this model the grain size distribution is such that mately to hydraulic conductivity by multiplying by 10.
most of the void space is hydraulically interconnected In shaly sandstones, however, there are two main
and almost the whole interstitial fluid is free to flow grainsize populations, one characterizing the sand
under an applied piezometric head. This inter- fraction and another the shaly portion. This generates
connected porosity may be referred to as an effective two porosity parameters: one effective hydraulic
hydraulic porosity (f e). The mean geometrical tortu- porosity, as for the clean sandstone, and a shale poros-
osity of the channels can be distinguished as a free ity whose pores are normally filled with bound water.
pore tortuosity having similar effects on both the elec- In the same way, we can distinguish two tortuosity
trical and hydraulic flows through the medium. For indices: one for the free water channels and another
such rocks analytical and experimental evidence for the bound water conduits. The first controls a large
support a relationship between the electrical and fraction of the electrical flow and the whole of the
hydraulic conductivities obtained from Archie’s and hydraulic transport through the medium. However,
Kozeny–Carman’s equations in the following form the second one, contributes only to the electrical trans-
(Croft, 1971; Kelly, 1977) port inside the shaly coatings. These two electrical
paths act essentially in a parallel association (de
k ˆ aF ⫺q ; …1† Lima and Sharma, 1990). Pure shales represent the
14 O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26

10 00 .0 10 00 .0

10 0.0 10 0.0
In trinsic P erm ea bility (m d )

In trinsic P erm ea bility (m d )


10 .0 1 0.0

1 .0 1.0

0.1 0.1
1.0 10 .0 1 00 .0 10 00 .0 0 .1 1.0 10 .0 10 0.0
R ea l P art o f C o nd uc tiv ity (X 0 .0 1 S /m ) Im a gina ry P art o f C o nd uctiv ity (X 10 -4 S /m )

Fig. 1. Relationship between intrinsic permeability and real and imaginary component of the electrical conductivities at 0.5 M NaCl electrolyte
salinity for typical shaly sandstones (data from Vinegar and Waxman (1984), 1984).

limits when fe ! 0; when one still has effective elec-  and the particular type of clay (b 1 is the charge
…rs ; a†
trical conduction but the medium may be almost density on a clay particle) and Qv is the clay compen-
impervious to water flows. sating ionic charge per unit pore volume of the
For shaly sandstone there is no straightforward sandstone (de Lima, 1995).
relationship between electrical and hydraulic conduc- Existing hydrophysical models have not used the
tivities. In Fig. 1 we plot the intrinsic permeability surface conduction effects and differences in tortuos-
data of Vinegar and Waxman (1984) for typical ities to estimate the total surface area contributed by
shaly sandstones, against their corresponding real shale/clay in an aquifer. At the same time results using
and imaginary electrical conductivity components. relations developed for clean formations remain ques-
From this figure we observe no obvious correlation tionable for use in shaly formations. This situation
between electrical and hydraulic parameters. requires that we widen the canvas by including
de Lima (1995), using a capillary approach for a other geoelectrical data in addition to resistivity
granular shaly coated model, has derived a semi- measurements to redeem this hopeless situation.
empirical equation relating the intrinsic permeability In the present work we review the petrophysical
of a porous medium to its porosity, formation factor basis for relating electrical and hydraulic parameters
and clay content as follows: of a granular material having dispersed charged parti-
" #q cles in its solid matrix. By reworking and expanding
fe F ⫺1 previous results for self-similar clay-coated models
k ˆ a0 ; …4†
1 ⫹ dQv described by de Lima and Sharma (1990, 1992) and
de Lima (1995) we developed a new permeability
where a 0 and q are also empirical constants, f e and F equation totally dependent on petrophysical para-
are the effective porosity and true formation factor of meters easily determined either from the surface elec-
the sandstone, d…ˆ rs =3ab1 † is a parameter controlled trical or well log measurements. For borehole logging
by the average sizes of the sand and clay particles the electrical resistivities measured with different
O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26 15

representation of this granular model outlining one


typical porous channel. Such a channel is character-
ized by wide pore bodies sequentially interconnected
through narrow pore throats.
The coating shaly structure also alters the effective
porosity and the electrical and hydraulic tortuosities
within the representative channel. An increase in the
clay proportion tends to reduce the effective free pore
diameter and to increase the electrical tortuosity of the
bound water conduits around the pore walls. These
will also introduce important effects on the hydraulic
conductivity of this porous medium. In the following
sections we will describe the basic electrical and
hydraulic characteristics of this granular shaly coated
Fig. 2. Sketch of a self-similar granular shaly sandstone model and sandstone.
one of its representative channel.

2.1. The electric effects


depths of investigation are required to find the neces-
sary parameters. From surface electrical measure- The electrical conductivity changes introduced by
ments the dc-resistivity and an induced polarization clays have been described by several authors
parameter are used for the conversion. Although the (Vacquier et al., 1957; Marshall and Madden, 1959;
IP phenomenon has been suggested to correlate with Vinegar and Waxman, 1984; de Lima and Sharma,
the aquifer permeability (Bodmer et al., 1968; Ogilvy 1990). The main enhancement is caused by the local
and Kuzmina, 1972; Worthington and Collar, 1984; ionic concentration gradients developed by membrane
Börner et al., 1996; Weller and Börner, 1996), to our filtration effects along constrictions on their pore
knowledge, this is the first report describing a firmly channels.
supported quantitative use of IP and resistivity for an Two characteristic lengths are distinguished in the
aquifer evaluation. assumed granular model (Fig. 2): a free pore length at
the grain-size scale and a pore throat constriction of
smaller dimension. An external electrical field will
2. The shaly sand model induce differentiated ionic fluxes inside this channel
that, along the constrictions, are opposed by the catio-
The electrical, hydraulic and other transport proper- nic selectivity of the clays. These filtration effects will
ties of shaly sandstones are strongly controlled by the change the local ionic distribution along the pore
geometrical arrangement of their grains, by the inter- channel. An ionic depletion is induced at the positive
facial processes occurring at boundaries between side of a constriction whereas a corresponding
water and clay particles, as well as by the saturation concentration is developed on its negative side.
level and the salinity of the interstitial water. These charge accumulations make the electrical
To simulate the physical behavior of shaly sand- conductivity of such sandstones frequency dependent,
stones we assume a self-similar porous model of described by a total complex current conductivity
spheroidal particles composed of ns insulating sand function (Dias, 1972; Klein and Sill, 1982; de Lima
grains of distributed sizes rs,i, and ni clay particles and Sharma, 1992). These are the main cause of the
also of distributed sizes ai and arranged essentially induced polarization effects observed in shaly sand-
as thin uniform coating structures around the sand stones (Vacquier et al., 1957; Marshall and Madden,
grains. The clays are electrically charged particles 1959).
that are compensated by exchangeable cations. Next, we will assume that for low frequency
Under a water wet condition these ions form diffuse ranges (such that s ⴱ q veⴱ ) the complex electri-
atmospheres around the clays. Fig. 2 is a schematic cal conductivity of a shaly sandstone is given by a
16 O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26

Bruggeman–Hanay type equation in the form (1992)



!m
2p
ⴱ m 1 ⫺ s cs =s w s csⴱ ˆ sⴱ ; …9†
s 0 ˆ s w fe ⴱ ; …5† 3 ⫺ p sh
1 ⫺ s cs =s 0ⴱ

where s sh is the complex conductivity of a shaly shell
where f e is the effective medium porosity and m and p is the clay volume fraction in the sandstone
a non-dimensional parameter that depends on the matrix. For spheroidal shaly coated particles we
average aspect ratio of the coated sand grains; s 0ⴱ may use Fricke’s result (Fricke, 1924) given as

and s cs are the complex conductivities of the
sandstone and of the clay-coated sand grains, wp
s cs ˆ s ; …10†
and s w the real conductivity of the electrolyte w ⫹ 1 ⫺ p sh
(de Lima and Sharma, 1992).
For low frequency ranges and for a dominant elec- where w is a shape factor. For small values of p …p ⬍
trical conduction through the porous electrolyte 0:3† Eqs. (9) or (10) allows to write
…s w q 兩s cs 兩† a binomial expansion on both the s cs;R ˆ gps sh;R ; …11†
numerator and denominator of Eq. (5) and retention
of only first order terms gives where g is also a geometrical constant ranging from
0.5 to 1. In cases where clays occurs as discrete parti-
1
s 0ⴱ ˆ ⴱ
‰s ⫹ m…F ⫺ 1†s cs Š; …6† cles dispersed through the sandstone matrix, de Lima
F w and Sharma (1990) also found that s cs ˆ ps sh : Thus,
where F is the usual Archie’s formation factor …F ˆ we have good support for using Eq. (11) to express p
f⫺m
e †: Eqs. (5) and (6), with appropriate substitutions,
in terms of the ratio between the effective grain
can also be applied to compacted clays and pure conductivity and the conductivity of its shaly coating.
shales. In this case, s 0ⴱ will represent s sh
ⴱ ⴱ
and s cs Further, for low porosity clays or shales and at low
will be replaced by the complex conductivity of frequency measurements, the following approxima-
wetted clay particles s cⴱ : tions hold (de Lima and Sharma, 1992)
The clay or the coated-clay conductivity will, in d1 …1 ⫹ d1 A†
general, be dependent on the conductivity of its inter- s sh;R ˆ s w; …12†
…1 ⫹ d1 A†2 ⫹ d21 B2
stitial electrolyte. However, as the effective porosity
of clay aggregates and shales is very small, the and
conductivity of their bond waters soon assume a
constant value independent of the water salinity of d21 B sw
the free pore electrolyte (Clavier et al., 1997). s sh;I ˆ ; …13†
…1 ⫹ d1 A†2 ⫹ d21 B2 v
Eq. (6) includes the following components:
where d1 ˆ b1 =aC1 ; b1 being the surface counterion
1
s 0;R ˆ ‰s w ⫹ m…F ⫺ 1†s cs;R Š; …7† density around a clay particle of radius a, and C1 is the
F cationic concentration in the electrolyte where the
coated spheres are immersed; A and B are parameters
and
dependent on frequency and the relaxation time of the
s 0;I ˆ m…1 ⫺ 1=F†s cs;I ; …8† charged clays. From Eqs. (12) and (13) we can write
d1 B
where the subscripts R and I stands for real and s sh;I ˆ s : …14†
v…1 ⫹ d1 A† sh;R
imaginary components of the complex conductivity
function. Combining Eqs. (14) and (9) we get
For a clay-coated spherical particle representative
of a shaly sandstone, and under the same low d1 B
s cs;I ˆ s : …15†
frequency range, we have from de Lima and Sharma v…1 ⫹ d1 A† cs;R
O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26 17

5E-4 5E-3
N=0.01 NaCl N=0.1 NaCl
Quadrature Conductivity (S/m)

Quadrature Conducutity (S/m)


2
y=0.01712*x+6.515E-05, r =0.9350
2.5E-4 3E-3

2
y=0.008225*x+4.141E-05, r =0.882

0 0
0 2E-2 4E-2 6E-2 8E-2 0 2E-2 4E-2 6E-2 8E-2
Matrix Conductivity (S/m) Matrix Conductivity (S/m)

Fig. 3. Imaginary component of the bulk conductivity of shaly sandstones versus the real component of its matrix conductivity for 0.01 and
0.1 M NaCl electrolytes (data from Vinegar and Waxman (1984), 1984).

Finally, substituting Eq. (15) in Eq. (8) we have dependent on the frequency and on the clay type in
m…1 ⫺ 1=F†d1 B the sandstone matrix.
s 0;I ˆ s cs;R ˆ lq s cs;R ; …16† Thus, by treating the electrical conductivity of a
v…1 ⫹ d1 A†
shaly sandstone as a complex parameter we derive
where l q is a non-dimensional parameter dependent explicit expressions relating the real and imaginary
on the geometrical features of the porous medium, on components of its bulk conductivity to the real compo-
the clay species present in the shaly shells and on the nent of the conductivity of its solid matrix. Now, we
frequency. have two alternative procedures to find (s cs,R) from
Eq. (16) shows that the imaginary component of the geoelectrical measurements: using Eq. (7) and the real
complex conductivity of a shaly sandstone depends component of conductivity measurements; and using
linearly on the real component of the volumetric Eq. (16) and the imaginary component of conductivity
conductivity of its solid matrix. The constant of measurements or an induced polarization parameter,
proportionality is a geometrical parameter slightly as we will show below.

5E -3 5E -3
N = 0.5 N aC l N = 1.0 N aC l
Im ag in a ry P a rt of C on du ctivity (S /m )

Im ag in a ry P a rt of C on du ctivity (S /m )

y= 0.021 14 *x+ 7.6 57 8E -05 , r 2 = 0.9 22 y= 0.017 34 *x+ 10 .54 E -0 5, r 2 =0 .93 9


2.5E -3 3 E -3

0 0
0 2E -2 4E -2 6 E -2 8E -2 0 2E -2 4E -2 6E -2 8E -2
M atrix C o nd uctiv ity (S /m ) M atrix C o ndu ctivity (S /m )

Fig. 4. Imaginary component of the bulk conductivity of shaly sandstones versus the real component of its matrix conductivity for 0.5 and 1 M
NaCl electrolytes (data from Vinegar and Waxman (1984), 1984).
18 O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26

In Figs. 3 and 4 we present the best correlation and Waxman, 1984)


results between the imaginary component of the p
2
conductivity of shaly sandstones measured by Vine- FE…v1 ; Av1 † ˆ ⫺ u…v1 A† ln A; …19†
gar and Waxman (1984) as a function of their grain p
volume conductivity. These were determined from with A being the ratio between the two used
straight-line fits to their bulk real component measure- frequencies and u ˆ tan⫺1 s 0;I =s 0;R is measured in
ments made with different water salinities. The figures milliradians.
show high levels of correlation (r 2 between 0.89 and As the IP phenomenon is a linear one the time
0.94) and that l q varies slightly from 0.00822 at domain and the frequency domain data are theoreti-
0.01 N to 0.01734 at 1 N. The fittings show very cally related by Fourier transforms. To a first approx-
small intercepts of about 4:1–10:5 × 10 ⫺5 S=m; imation we may use in practice the following result
attesting the general validity of Eq. (16). relating the peak chargeability M(0) to the maximum
The detection of IP effects in the time domain is FE effect (Zonge et al., 1972)
made by measuring both the maximum voltage during
M…0†
the charging period of a dc-current flow, and the tran- FE ˆ : …20†
1 ⫺ M…0†
sient decay of the electrical potential as a function of
time, after the current is turned off. The maximum
voltage is used to compute an apparent resistivity 2.2. The hydraulic effect
function (r a). The normalized integration of the
decaying voltage at a specified time window gives The intrinsic permeability of a porous medium
an apparent, time-dependent, chargeability (Ma) for (dimension L 2) is a measure of its frictional resistance
that window as to a single-phase fluid flowing through it. It depends
on the pore size distribution, on the roughness and
1 Zt2 constrictions of pore space and on the tortuosity and
Ma …t† ˆ V dt: …17†
V0 Dt t1 IP connectivity of their internal pore channels. The most
extensively used approach to simulate the hydraulic
where Dt ˆ t2 ⫺ t1 is the width of the window and V0 behavior of granular rocks is through the Kozeny–
is the maximum voltage observed during the charging Carman theory. In this approach, the granular porous
period. With modern instrumentation, the usual medium is statistically modeled as a spatial array of
practice is to sample the decaying voltage through a capillary channels having different lengths and cross-
sequence of windows. In each window a partial appar- sections. Theoretical and experimental results have
ent chargeability is measured. These values are then attested the practical validity of such approximation
used to compute an weighted average chargeability (Wyllie and Spangler, 1952; Terzaghi, 1955; Haring
for the decaying period. and Greenkorn, 1970; Scheidegger, 1974; Brace,
In the frequency domain both the amplitude and the 1977; Wong et al., 1984; Clennell, 1997).
phase shift of the apparent impedance of the earth is Following de Lima (1995) we start from the
measured using different frequencies. A frequency Kozeny–Carman equation, derived for a porous
effect (FE) can be evaluated by comparing the ampli- medium modeled as a bundle of sinuous capillary
tude measurements made at a high alternating channels, given as
frequency with the one from a much lower frequency  
VP q
taken as a dc value k ˆ afe ; …21†
SA t h
兩rⴱ …v1 †兩 ⫺ 兩rⴱ …v2 †兩 where a is a dimensionless shape factor related to the
FE ˆ ; …18†
兩rⴱ …v2 †兩 statistical average of the channel perimeter to its
cross-sectional area, f e is the effective porosity, SA/
with v1 p v2 : An alternative procedure uses the VP is the specific surface area referred to the effective
phase shift measurement made at an intermediate pore volume, and t h is an average hydraulic tortuosity
frequency as (Shuey and Johnson, 1973; Vinegar for the medium. The constant a for nearly uniform
O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26 19

pore size distributions is about 0.5 (Carman, 1956). granular model is


The fractional exponent q in Eq. (21) is required to "n #
Xs X
nc
take into account the fractal nature of the pore space DAs;c ˆ 4p rs;i ⫹
2
ai :
2
…26†
of a real porous rock. iˆ1 iˆ1
From Fig. 2 we assume, as was performed by de
With p given as
Lima (1995), that the hydraulic tortuosity of such a
model can be approximated by the mean electric p ˆ 1 ⫺ DVs =DVcs ;
tortuosity of the free pore channels, given as and
tˆ f…1⫺m†
e : …22† DVc ˆ p…1 ⫺ fsh †DVcs

The conductive contribution due to the shaly shells we may write the following relations:
is described by a different tortuosity referred as the X
ns
3
X
nc

bound-water tortuosity tb ˆ …1 ⫺ fe †=m…1 ⫺ fm rs;i a3i


e † (de X
ns
iˆ1 iˆ1
Lima and Sharma, 1992). Eq. (22) has been derived 3
rcs;i ˆ ˆ : …27†
for clean sandstones, by combining the definitions of iˆ1
1⫺p p…1 ⫺ fsh †
Archie’s formation factor and Carman’s hydraulic Using the above relations, the specific surface area
tortuosity (Wyllie and Rose, 1950; Winsauer et al., related to the effective pore volume of the sample
1952). The similarity between hydraulic and electric (dimension L ⫺1) can be written as
tortuosities has been also pointed out by Clennell  
S r
(1997) and many others. A similar approach has Ss;c ˆ A ˆ Ss 1 ⫹ p…1 ⫺ fsh † s ; …28†
been implicitly used by Mualem and Friedman VP a
(1991) and Weerts et al. (1999) for modeling the where Ss ˆ 3…1 ⫺ fe †=fe rcs is the specific surface
bulk soil electrical conductivity of saturated and area of a clean sandstone of the same texture and
unsaturated soils. porosity, and rcs ; rs and a are the harmonic averages
The specific surface for the model may be for the coated sands, the sand cores and the clay sizes,
computed as follows. A representative sample of respectively. Eq. (28) has been obtained on the
volume DV0 of the model contains ns sand grains assumption of small values of p and f sh such that
and nc clay particles. Two classes of pore space products containing the square of these quantities
exist: one, of dimension in the sand-size scale, contri- were neglected. For many shaly sandstones, there is
butes to the effective porosity, f e; another, within a a direct relation between their matrix electrical
clay-size dimension, characterizes the shale porosity, conductivity, the shale proportion and the shale
f sh. If p is the shale volume fraction, the total conductivity, allowing to rewrite Eq. (28) as
porosity, f , for this sample is
Ss;c ˆ Ss …1 ⫹ dc s cs;R †; …29†
f ˆ fe ⫹ pfsh …1 ⫺ fe †: …23†
where dc ˆ …1 ⫺ fsh †rs =a gs sh depends on the proper-
The effective void volume in the sample is ties of the clay components and the average size of the
sand fraction.
fe For montmorillonite and kaolinite gels in distilled
DVvcs ˆ DVcs ; …24†
1 ⫺ fe water, experimental data from Cremers and Laudelout
(1966) allow us to find the minimum values for the
where DVcs is the volume fraction occupied by the electrical conductivity for these compacted wet clays
shaly coated matrix. For spherical particles as s cmont ˆ 0:703 S/m and s ckao ˆ 0:083 S=m (de
Lima and Sharma, 1990). For a critical conductivity
4 fe X ns
of the electrolyte saturating the shaly shells of
DVcs ˆ p r3 : …25†
3 1 ⫺ fe iˆ1 cs;i 2.5 S/m, a shaly porosity in the range of 20–30%,
and an averaged representative grain size ratio rs =a ˆ
The total solid surface area exposed to fluid in this 200 we find d c varying from 30 to 150 m/S,
20 O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26

1 00 00 0.0

100 00.0
In trin sic P erm e ab ility (m d)

1 00 0.0

q =2 .0

100 .0

q =2 .5
1 0.0

1 .0

M on tm orillon ite
K a o lin ite
0 .1
0 .00 1 0 .0 10 0.10 0 0 .50 0
L ith op oro sity Fa ctor

Fig. 5. Theoretical intrinsic permeability versus the lithoporosity factor for typical shaly sandstone models.

respectively, for montmorillonite and kaolinite size distributions are such that we have two distinct
aggregates or shales. values of d c equal to 50 and 200 m/S. In Fig. 5 we plot
Combining the results expressed by Eqs. (22) and in a log–log scale the modeled permeability values
(29) into Eq. (21) we obtain against a lithoporosity factor for shaly sandstones
" …m⫺1⫹1=q† #q defined as Lfe ˆ fe…m⫺1⫹1=q† =…1 ⫹ dc s cs †: We can see
fe from the figure that the effect of a change in d c is to
k ˆ a0 ; …30†
1 ⫹ dc s cs displace the span of the permeability–lithoporosity
relationship. The exponent q determines the inclina-
where a0 ˆ a=Sqs has a topological dimension L 2, d c is tion of the straight line segments, whereas a 0 defines
a lithologic parameter dependent on the particle size the starting point at s cs ! 0 and fe ! 1:
distributions for sands and clays, and the exponent q is
expected to depend on the coated particle shapes and 2.3. Experimental results
the packing structure of the sandstone. In general, a 0
varies from about 180,000 mm 2 in fine grained …r cs ˆ The permeability Eq. (30) has been tested with
100 mm† to 50,000 mm 2 in medium grained …rcs ˆ experimental core data published by Vinegar and
300 mm† shaly sandstones and q, the space filling Waxman (1984). This data set includes sandstone
fractal dimension, is in the range from 2 to 3. samples representative of different US oil fields. The
To show the effect of changes on the characteristic measurements were made under controlled condi-
parameters of Eq. (30) we constructed Fig. 5. Starting tions, at five different NaCl electrolyte resistivities,
from a clean sandstone of initial effective porosity of covering a range from 0.07 to 8.5 V m. Thus, they
35% we add, in steps, small amounts of clays up to a include both saline and fresh water electrolytes.
maximum p ˆ 20%; ending with a shaly sandstone of Values of s cs and F were determined from linear
fe ˆ 18:75%: Two types of clays are specified by best fitting the real component of their electrical
…1† …2†
s sh ˆ 0:078 S=m and s sh ˆ 0:685 S=m: The grain conductivity measurements. The effective porosity
O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26 21

1 00 0

1 00
P erm e ab ility (m d)

10

1
0 .01 0 0 .10 0 1 .0 00
L ith op oro sity fa cto L e

Fig. 6. Relationship between the intrinsic permeability and lithoporosity factor for shaly sandstone (data from Vinegar and Waxman (1984).

f e of each sample has been computed from the F permeability of the São Sebastião aquifer using well
values using an average m for the data set that log and surface electrical geophysical measurements.
includes only the samples for which the intrinsic The São Sebastião Formation is one of the most
permeability is above 1 md. As the clays in the important aquifer in the Recôncavo Basin, Bahia—
samples were described as mixtures of kaolinite, illite Brazil. It is composed of thick shaly sandstones
and montmorillonite we estimated d c to be between 50 having interlayered shales and siltstones. The sand-
and 100 m/S. Eq. (30) was fitted to their experimental stones were formed as stacked channel and bar depos-
data by plotting log(k) versus log…L fe † as shown in its of a large braided alluvial environment. The shales
Fig. 6 for dc ˆ 100 m=S: The best fit linear routine and siltstones are lacustrine deposits related to rapid
gives a0 ˆ 180; 015 md and q ˆ 2:78 with a regres- flooding episodes within a tectonically active intra-
sion coefficient r2 ˆ 0:9054: Two other values for d c continental basin (de Lima, 1993). Fig. 7 shows the
were used with the same data. For dc ˆ 50 we find statistical distribution of porosities and permeabilities
a0 ˆ 299; 610 md, q ˆ 3:51 and r2 ˆ 0:9183: For measured on 53 sandstones samples selected from
dc ˆ 200 m=S we obtain a0 ˆ 181; 920 md, q ˆ drill cores of a deep well drilled to a depth of
2:35 and r 2 ˆ 0:8895: From these results we find the 1025 m. The measured porosities range from 18.8 to
data can be satisfactorily described by assuming 34.9%, with a median of 25.7%. Their permeabilities
dc ˆ 100: cover a wide spectrum having a bimodal logarithmic
Although the adjusted parameters are naturally not distribution ranging from less than 100 to 2750 md,
universal, they may be taken as characteristics for a with a median of 794 md.
wide class of shaly sandstones. If so, they can be used Fig. 8 shows the geophysical log of a water produc-
for initial estimates of the order of magnitude of the tion well drilled in the São Sebastião Formation in
intrinsic permeability of shaly formations. In this way, Recôncavo Basin, Bahia, Brazil. The log suite
the determined parameters were applied in this work consists of: (i) the deep induction and the short normal
as first approximations to estimate the intrinsic resistivity curves responding to both the virgin
22 O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26

Fig. 7. Frequency distribution of porosity and permeability for shaly


sandstones of the São Sebastião Formation.

conditions and the mud filtrate invaded zone in the Fig. 8. Porosity, clay content and intrinsic permeability estimated
reservoir rock; (ii) the self-potential log and the from well log measurements on a water well drilled in Bahia, Brazil.
natural gamma ray curve; and (iii) the sonic log
converted to a total porosity curve. The resisitivity
of the groundwater was measured directly on samples compacted and the sonic data requires corrections
…rw ˆ 70 V m†; whereas that of the mud filtrate was before the application of Wyllie formula (Jorden and
estimated from the mud resistivity and from the SP Campbell, 1986). Permeability values were computed
deflection …r mf ˆ 30:3 V m†: with Eq. (30) using a 0 and q as determined for the
Both s cs and F were determined using Eq. (7) sandstone samples (Vinegar and Waxman, 1984). In
written for the flushed and the virgin conditions, general, the computed k are in a range from 100 to
respectively. The clay volume fraction p was 2200 md with the more frequent values laying
computed from Eq. (9) with s sh ˆ 0:085 S=m a between 1000 and 2000 md. This range compares
value close to that of a kaolinitic aggregate. The favorably with core measurements on true sandstone
computed p is in good agreement with Vcl estimated samples as described above.
from the gamma ray log using a conventional empiri- In Fig. 9 we show a time-domain Schlumberger IP-
cal equation (Asquith, 1990). Effective porosities resistivity sounding made over the São Sebastião
were derived from the formation resistivity factor Formation using a SYSCAL R2 resistivity-meter.
using an average m equal to 1.83 (Börner et al., The apparent resistivity data were inverted using a
1993). The sonic porosities are too high as compared non-linear least square procedure based on horizon-
to f e, suggesting the sandstones are weakly tally layered earth model (Koefoed, 1979). As shown
O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26 23

Fig. 9. Geoelectrical sounding data used in the petrophysical interpretation of shaly sandstone aquifer in Bahia, Brazil.

on the upper part of the figure the experimental data two next layers with resistivities of 730 and
has been satisfactorily fitted to a five layers earth 1320 V m. Below 116 m the aquifer is underlain by
model. The two upper layers are in the top soil a sequence dominated by shales (20 V m).
above the water table. The main aquifer includes the Simultaneously, partial apparent chargeabilities
24 O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26

were measured at four IP windows centered at 120, by de Lima and Sharma (1990, 1992) to describe the
220, 420 and 820 ms, as depicted in the bottom of electrical properties of shaly sandstones. We have
Fig. 9. These data sets were interpreted using the shown both theoretically and with experimental
quasi-linear geoelectrical approximation developed data that the electrical resistivity and the induced
by Patella (1973). In his work Patella introduced a polarization parameters carry direct information
ficticious apparent resistivity …r 0a † for the Schlumberger about the volume conductivity of the sandstone
array expressed as matrix. Thus, the well log measurements obtained
with tools having different depths of investigation,
r 0a ˆ Ma ra ; …31† as well as surface IP-resistivity measurements can
where r a is the usual apparent resistivity function. The be used to find F, s cs and p for shaly sandstone
analytic expression he found for r 0a is formally formations. Practical application of these procedures
equivalent to that of the conventional r a. Thus, shows good agreement with estimates based on
because r 0i ˆ Mi ri ; we can use the same iterative different approaches and with experimental measure-
procedure for inversion of r a to invert the apparent ments on samples.
chargeability data. We have also developed a new procedure for
The model geometry was fixed according to the predicting the permeability of sandstones based on
apparent resistivity interpretation, and the inversion petrophysical parameter readily extractable from elec-
scheme was applied to find the partial Mi’s for the trical logs and surface geoelectrical measurements.
different layers. The best-fit chargeability models We derive a Kozeny–Carman type equation from
are shown on the right bottom of Fig. 9. Then, these first principles, showing the remarkable influence of
partial values were used to obtain a close estimate for a lithoporosity factor—the combination of a geome-
the intrinsic chargeability of the layers. trical porosity–tortuosity term, with a length pore
For each layer the partial chargeabilities were scale, clay type and volume content in the sandstone
approximately described by an exponential function matrix, as expressed by the factor d cs cs. In this treat-
M…t† ˆ M0 exp…⫺t=t†; where M0 can be taken as the ment the hydraulic tortuosity is determined as the
peak chargeability and t as an average relaxation time average electrical tortuosity along the free pore
for the induced polarization in the layer. By this way electrolyte. The equation was firmly tested with
we find for the two aquifer layers M0 …3† ˆ 8:7 mV=V experimental data measured on different shaly sand-
and M0 …4† ˆ 5:4 mV=V: These results were converted stone samples published by Vinegar and Waxman
to FE’s and then to phase shifts using Eqs. (20) and (1984). Next, it was applied as a first approximation
(19), the latter written assuming a 100 times to estimate sandstone permeabilities from electrical
frequency variation. The true bulk resistivities and logs and surface geoelectrical sounding measure-
the u values of the layers allows to find their s 0,I’s. ments. The inferred values compare consistently
From Eq. (16) we found the s cs,R’s using l q ˆ with the sample measurements. These estimated
0:0061; a value extrapolated from the data of Figs. 3 values of permability can easily be converted to that
and 4, for a NaCl electrolyte of 0.001 M salinity of hydraulic conductivity to be used for groundwater
…rw ˆ 70 V m†: Finally using Eq. (30) together with management.
the following petrophysical parameters for the
sandstones f e ˆ 25%; m ˆ 1:83; dc ˆ 100 m=S; and
a0 ˆ 180; 015 md and q ˆ 2:78; as found for the Acknowledgements
Vinegar and Waxman data, we obtain k3 ˆ 1500 md
and k4 ˆ 1750 md for the permeabilities of these two We wish to thank the Brazilian agencies Finan-
sandstone bodies. ciadora de Estudos e Projetos (FINEP) and the
National Council for Scientific Development
(CNPq) for providing the financial support for
3. Conclusions this work. We would like to thank H. Sato and
B. Clennell for the critical revision of this
We briefly review and expand the theory proposed manuscript.
O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26 25

References Acoustic Logging. Soc. Petr. Eng., Mono. 10 H. L. Doherty


Series.
Archie, G.E., 1950. Introduction to petrophysics of resevoir rocks. Katz, A.J., Thompson, A.H., 1985. Fractal sandstone pores: impli-
Bull. AAPG 34, 943–961. cations for conductivity and pore formation. Phys. Rev. Lett. 54,
Asquith, G., 1990. Log evaluation of shaly sandstones: a practical 1325–1328.
guide. Am. Assoc. Petr. Geol., Course Notes No. 31. Kelly, W.E., 1977. Geoelectric sounding for estimating hydraulic
Bodmer, R., Ward, S.H., Morrison, H.F., 1968. On induced conductivity. Ground Water 15, 420–425.
electrical polarization and groundwater. Geophysics 33, 805– Klein, J.D., Sill, W.R., 1982. Electrical properties of artificial
821. clay-bearing sandstone. Geophysics 47, 1593–1605.
Börner, F., Gruhne, M., Schön, J., 1993. Contamination indications Koefoed, O., 1979. Geosounding Principles 1—Resistivity Sound-
derived from electrical properties in the low frequency range. ing Measurements, Elsevier, Amsterdam.
Geophys. Prosp. 41, 83–89. Korvin, G, 1992. Fractal Models in Earth Sciences, Elsevier,
Börner, F.D., Schopper, J.R., Weller, A., 1996. Evaluation of trans- Amsterdam.
port and storage properties in the soil and groundwater zone Kosinski, W.K., Kelly, W.E., 1981. Geoelectric soundings for
from induced polarization measurements. Geophys. Prosp. 44, predicting aquifer properties. Ground Water 19, 163–171.
583–601. Marshall, D.J., Madden, T.R., 1959. Induced polarization, a study of
Brace, W.F., 1977. Permeability from resistivity and pore shape. its causes. Geophysics 24, 790–816.
J. Geophys. Res. 82, 3343–3349. Mualem, Y., Friedman, S.P., 1991. Theoretical prediction of
Carman, P.C., 1956. Flow of Gases Through Porous Media, electrical conductivity in saturated and unsaturated soil. Water
Butterworth, London. Resour. Res. 27, 2771–2777.
Clennell, B., 1997. Tortuosity: a guide through the maze Special Ogilvy, A.A., Kuzmina, E.N., 1972. Hydrogeologic and engineer-
Publication 1997. In: Lovell, M.A., Harvey, P.K. (Eds.). ing-geologic possibilities for employing the method of induced
Developments in Petrophysics, Geological Society, London, potencials. Geophysics 37, 839–861.
p. 122. Patella, D., 1973. new parameter for the interpretation of induced
Cremers, A.E., Laudelout, H., 1966. Surface mobilities of cations in polarization field prospecting (time-domain). Geophys. Prosp.
clays. Proc. Soil Sci. Soc. Am. 30, 570–576. 21, 315–329.
Croft, M.G., 1971. method of calculating permeability from electric Scheidegger, A.E., 1974. The physics of flow through porous
logs. US Geol. Surv., Prof. Paper 750, 265–269. Media. 3rd ed. University of Toronto Press.
Clavier, C., Coates, G., Dumanoir, J., 1997. The theoretical and Shuey, R.T., Johnson, M., 1973. On the phenomenology of electri-
experimental bases for the “dual water” model for the interpre- cal relaxation in rocks. Goephysics 38, 37–48.
tation of shaly sands. 52nd Ann. Fall Tech. Conf., Soc. Petr. Scott, T.D., C. Purdy, C., 1988. Log derived permeability and
Engr., Paper SPE. 2961. flushed zone saturation utilising high frequency electromagnetic
de Lima, O.A.L., Sharma, M.M., 1990. A grain conductivity wave attenuation and propagation time. 29th Ann. Log. Symp.
approach to shaly sands. Geophysics 50, 1347–1356. Trans., Soc. Prof. Well Log Analysts, Paper I.
de Lima, O.A.L., Sharma, M.M., 1992. A generalized Maxwell– Sen, P.N., Straley, C., Kenion, W.E., Whittingham, M.S., 1990.
Wagner theory for membrane polarization in shaly sands. Surface-to-volume ratio, charge density, nuclear magnetic
Geophysics 57, 431–440. relaxation, and permeability in clay-bearing sandstones.
de Lima, O.A.L., 1993. Geophysical evaluation of sandstone Geophysics 55, 61–69.
aquifers in the Recôncavo-Tucano Basin, Bahia—Brazil. Sighal, D.C., Sri Niwas, Shaked, M., Adam, E.M., 1998. Estimation
Geophysics 58, 1689–1702. of hydraulic characteristics of alluvial aquifers from electrical
de Lima, O.A.L., 1995. Water saturation and permeability from resistivity data. J. Geol. Soc. India 51, 165–183.
resistivity, dielectric and porosity logs. Geophysics 60, 1756– Sri Niwas, Singhal, D.C., 1981. Estimation of aquifer transmissivity
1764. from Dar-Zarrouk parameters in porous media. Hydrol. J. 50,
Dias, C.A., 1972. Analytical model for a polarizable medium at 393–399.
radio and lower frequencies. J. Geophys. Res. 77, 4945–4956. Sri Niwas, Singhal, D.C., 1985. Aquifer transmissivity of porous
Fricke, H., 1924. mathematical treatment of the electrical con- media from resistivity data. Hydrol. J. 82, 143–153.
ductivity and capacity of disperse system. Phys. Rev. 24, Terzaghi, K.V., 1955. Influence of geological factors on the
575–587. engineering properties of sediments. Econ. Geol. 15, 557–
Haring, R.E., Greenkorn, R.A., 1970. statistical model of a porous 618.
medium with nonuniform pores. Am. Inst. Chem. Eng. J. 16, Turcotte, D.L., 1992. Fractals and Chaos in Geology and Geophy-
477–483. sics, Cambridge University Press, New York.
Heigold, P.C., Gilkeson, R.H., Castwright, K., Reed, P.C., 1980. Vacquier, V., Holmes, C.R., Kintzinger, P.R., Lavergne, M., 1957.
Aquifer transmissivity from surficial electrical measurements. Prospecting for ground water by induced electrical polarization.
Groundwater 17, 330–345. Geophysics 22, 660–687.
Hubbert, M.K., 1940. The theory of groundwater motions. J. Geol- Vinegar, H.J., Waxman, M.H., 1984. Induced polarization of shaly
ogy 48, 785–944. sands. Geophysics 49, 1267–1287.
Jorden, J.R., Campbell, F.L., 1986. Well Logging II—Electric and Yadav, G.S., Abolfazli, H., 1998. Geoelectrical soundings and their
26 O.A.L. de Lima, Sri Niwas / Journal of Hydrology 235 (2000) 12–26

relationship to hydraulic parameters in semiarid regions of Wong, P., Koplik, J., Tomanic, J.P., 1984. Conductivity and perme-
Jalore, Northwestern India. J. Appl. Geophys. 39, 35–51. ability of rocks. Phys. Rev. B 30, 6604–6614.
Zonge, K.L., Sauck, W.A., Sumner, J.S., 1972. Comparison of time, Worthington, P.F., Collar, F.A., 1984. The relevance of induced
frequency, and phase measurements in induced polarization. polarization to quantitative formation evaluation. Marine and
Geophys. Prosp. 20, 626–648. Petr. Geol. 1, 14–26.
Weerts, A.H., Bouten, W., Verstraten, J.M., 1999. Simultaneous Wyllie, M.R.J., Rose, W.D., 1950. Some theoretical considerations
measurement of water retention in soil: testing the Mualem– related to quantitative evaluation of the physical characteristics
Friedman tortuosity model. Water Resour. Res. 35, 1781–1787. of reservoir rock from electric log data. Trans. Am. Inst. Mech.
Weller, A., Börner, F.D., 1996. Measurements of spectral induced Engng 189, 105–118.
polarization for environmental purposes. Environ. Geol. 27, Wyllie, M.R.J., Spangler, M.B., 1952. Application of electrical
329–334. resistivity measurements to problems of fluid flow in porous
Winsauer, W.O., Shearin, H.M., Masson, P.H., Williams, M., 1952. media. Bull. Am. Assoc. Petr. Geol. 36, 359–403.
Resistivity of brine-saturated sands in relation to pore geometry.
Bull. Am. Ass. Pet. Geol. 36, 253–277.

You might also like